Skip to main content
Log in

Comparison between direct numerical simulations and effective models for fluid-porous flows using penalization

  • Published:
Meccanica Aims and scope Submit manuscript

Abstract

This work is devoted to a numerical study of two-dimensional incompressible flows in a fluid-porous medium system placed on an impermeable wall. Flow in the whole system is studied either at the pore scale or at the Darcy scale (using a one-domain approach) and the models at both scales are solved with a penalization method using the same formal Navier–Stokes equations modified by a Darcy-like term. Several effective medium penalized models are considered for the simulations. Various flow regimes are investigated ranging from laminar to turbulent. The velocity profiles inside and outside the porous medium show significant discrepancies between the different penalization models compared to the direct numerical simulations. This work motivates further studies about the spatial variations of the penalization coefficient to be introduced in the effective medium models in order to better reproduce the physics near the fluid-porous medium boundaries.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. DNS results are available from the authors upon request.

References

  1. Ahmed SR, Ramm G, Faltin G (1984) Some salient features of the time-averaged ground vehicle wake. In SAE technical paper, vol 02 . SAE International. https://doi.org/10.4271/840300. https://doi.org/10.4271/840300

  2. Angot Ph, Bruneau CH, Fabrie P (1999) A penalization method to take into account obstacles in incompressible viscous flows. Numer Math 81(4):497–520

    Article  MathSciNet  Google Scholar 

  3. Ardekani MN, Al Asmar L, Picano F, Brandt L (2018) Numerical study of heat transfer in laminar and turbulent pipe flow with finite-size spherical particles. Int J Heat Fluid Flow 71:189–199. https://doi.org/10.1016/j.ijheatfluidflow.2018.04.002

    Article  Google Scholar 

  4. Arquis E, Caltagirone JP (1984) Sur les conditions hydrodynamiques au voisinage d’une interface milieu fluide-milieux poreux: application à la convection naturelle. Comptes Rendus de l’Academie des Sciences 299:1–4

    Google Scholar 

  5. Auriault J-L (2009) On the domain of validity of Brinkman’s equation. Transp Porous Media 79(2):215–223. https://doi.org/10.1007/s11242-008-9308-7

    Article  MathSciNet  Google Scholar 

  6. Beavers GS, Joseph DD (1967) Boundary conditions at a naturally permeable wall. J Fluid Mech 30(01):197. https://doi.org/10.1017/s0022112067001375

    Article  ADS  Google Scholar 

  7. Bergmann M, Iollo A (2016) Bioinspired swimming simulations. J Comput Phys 323:310–321

    Article  ADS  MathSciNet  Google Scholar 

  8. Breugem WP, Boersma BJ (2005) Direct numerical simulations of turbulent flow over a permeable wall using a direct and a continuum approach. Phys Fluids 17(2):025103. https://doi.org/10.1063/1.1835771

    Article  ADS  MATH  Google Scholar 

  9. Breugem WP, Boersma BJ, Uittenbogaard RE (2006) The influence of wall permeability on turbulent channel flow. J Fluid Mech 562:35. https://doi.org/10.1017/s0022112006000887

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Breugem W-P, van Dijk V, Delfos R (2014) Flows through real porous media: X-ray computed tomography, experiments, and numerical simulations. J Fluids Eng. https://doi.org/10.1115/1.4025311

    Article  Google Scholar 

  11. Bruneau C-H (2000) Boundary conditions on artificial frontiers for incompressible and compressible Navier–Stokes equations. M2AN 34(2):303–314

    Article  MathSciNet  Google Scholar 

  12. Bruneau C-H, Fabrie P (1996) New efficient boundary conditions for incompressible Navier–Stokes equations : a well-posedness result. M2AN 30(7):815–840

    Article  MathSciNet  Google Scholar 

  13. Bruneau C-H, Khadra K (2016) Highly parallel computing of a multigrid solver for 3d Navier–Stokes equations. J Comput Sci 17(1):35–46

    Article  MathSciNet  Google Scholar 

  14. Bruneau C-H, Saad M (2006) The 2d lid-driven cavity problem revisited. Comput Fluids 35(3):326–348

    Article  Google Scholar 

  15. Bruneau C-H, Tancogne S (2018) Far field boundary conditions for incompressible flows computation. J Appl Anal Comput 8(3):690–709

    MathSciNet  Google Scholar 

  16. Cieszko M, Kubik J (1999) Derivation of matching conditions at the contact surface between fluid-saturated porous solid and bulk fluid. Transp Porous Media 34(1/3):319–336. https://doi.org/10.1023/a:1006590215455

    Article  Google Scholar 

  17. Ergun S (1952) Fluid flow through packed columns. Chem Eng Prog 48:48–89

    Google Scholar 

  18. Hardy B, De Wilde J, Winckelmans G (2019) A penalization method for the simulation of weakly compressible reacting gas-particle flows with general boundary conditions. Comput Fluids 190:294–307

    Article  MathSciNet  Google Scholar 

  19. Isoardi L, Chiavassa G, Ciraolo G, Haldenwang P, Serre E, Ghendrih Ph, Sarazin Y, Schwander F, Tamain P (2010) Penalization modeling of a limiter in the tokamak edge plasma. J Comput Phys 229(6):2220–2235

    Article  ADS  MathSciNet  Google Scholar 

  20. Jäger W, Mikelić A (2009) Modeling effective interface laws for transport phenomena between an unconfined fluid and a porous medium using homogenization. Transp Porous Media 78(3):489–508. https://doi.org/10.1007/s11242-009-9354-9

    Article  MathSciNet  Google Scholar 

  21. Jamet D, Chandesris M, Goyeau B (2008) On the equivalence of the discontinuous one- and two-domain approaches for the modeling of transport phenomena at a fluid/porous interface. Transp Porous Media 78(3):403–418. https://doi.org/10.1007/s11242-008-9314-9

    Article  MathSciNet  Google Scholar 

  22. Kajishima T, Takiguchi S, Hamasaki H, Miyake Y (2001) Turbulence structure of particle-laden flow in a vertical plane channel due to vortex shedding. JSME Int J Ser B 44(4):526–535. https://doi.org/10.1299/jsmeb.44.526

    Article  Google Scholar 

  23. Keetels GH, D’Ortona U, Kramer W, Clercx HJH, Schneider K, van Heijst GJF (2007) Fourier spectral and wavelet solvers for the incompressible Navier–Stokes equations with volume-penalization: convergence of a dipole-wall collision. J Comput Phys 227(2):919–945

    Article  ADS  MathSciNet  Google Scholar 

  24. Kevlahan NK-R, Ghidaglia J-M (2001) Computation of turbulent flow past an array of cylinders using a spectral method with Brinkman penalization. Eur J Mech B Fluids 20(3):333–350

    Article  Google Scholar 

  25. Khadra K, Angot P, Parneix S, Caltagirone J-P (2000) Fictitious domain approach for numerical modelling of Navier–Stokes equations. Int J Numer Methods Fluids 34(8):651–684. https://doi.org/10.1002/1097-0363(20001230)34:8<651::aid-fld61>3.0.co;2-d

    Article  MATH  Google Scholar 

  26. Larson RE, Higdon JJL (1986) Microscopic flow near the surface of two-dimensional porous media. Part 1. Axial flow. J Fluid Mech 166(–1):449. https://doi.org/10.1017/s0022112086000228

    Article  ADS  MATH  Google Scholar 

  27. Larson RE, Higdon JJL (1987) Microscopic flow near the surface of two-dimensional porous media. Part 2. Transverse flow. J Fluid Mech 178(–1):119. https://doi.org/10.1017/s0022112087001149

    Article  ADS  MATH  Google Scholar 

  28. Lasseux D, Abbasian-Arani AA, Ahmadi A (2011) On the stationary macroscopic inertial effects for one phase flow in ordered and disordered porous media. Phys Fluids 23:073–103

    Article  Google Scholar 

  29. Lasseux D, Valdés-Parada FJ, Bellet F (2019) Macroscopic model for unsteady flow in porous media. J Fluid Mech 862:283–311. https://doi.org/10.1017/jfm.2018.878

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. Mimeau C, Cottet G-H, Mortazavi I (2016) Direct numerical simulations of three-dimensional flows past obstacles with a vortex penalization method. Comput Fluids 136:331–347

    Article  MathSciNet  Google Scholar 

  31. Mittal R, Iaccarino G (2005) Immersed boundary methods. Annu Rev Fluid Mech 37(1):239–261. https://doi.org/10.1146/annurev.fluid.37.061903.175743

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Neale G, Nader W (1974) Practical significance of Brinkman’s extension of Darcy’s law: coupled parallel flows within a channel and a bounding porous medium. Can J Chem Eng 52(4):475–478. https://doi.org/10.1002/cjce.5450520407

    Article  Google Scholar 

  33. Ochoa-Tapia JA, Whitaker S (1998) Momentum jump condition at the boundary between a porous medium and a homogeneous fluid: inertial effects. J Porous Media 1(3):201–217

    MATH  Google Scholar 

  34. Ochoa-Tapia JA, Whitaker S (1995) Momentum transfer at the boundary between a porous medium and a homogeneous fluid—I. Theoretical development. Int J Heat Mass Transf 38(14):2635–2646. https://doi.org/10.1016/0017-9310(94)00346-w

    Article  MATH  Google Scholar 

  35. Paéz-García CT, Valdés-Parada FJ, Lasseux D (2017) Macroscopic momentum and mechanical energy equations for incompressible single-phase flow in porous media. Phys Rev E. https://doi.org/10.1103/physreve.95.023101

    Article  Google Scholar 

  36. Saffman PG (1971) On the boundary condition at the surface of a porous medium. Stud Appl Math 50(2):93–101. https://doi.org/10.1002/sapm197150293

    Article  MathSciNet  MATH  Google Scholar 

  37. Samanta A, Vinuesa R, Lashgari I, Schlatter P, Brandt L (2015) Enhanced secondary motion of the turbulent flow through a porous square duct. J Fluid Mech 784:681–693. https://doi.org/10.1017/jfm.2015.623

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Schneider K, Neffaa S, Bos WJT (2011) A pseudo-spectral method with volume penalisation for magnetohydrodynamic turbulence in confined domains. Comput Phys Commun 182(1):2–7

    Article  ADS  MathSciNet  Google Scholar 

  39. Shirokoff D, Nave J-C (2015) A sharp-interface active penalty method for the incompressible Navier–Stokes equations. J Sci Comput 62(1):53–77

    Article  MathSciNet  Google Scholar 

  40. Spietz HJ, Hejlesen MM, Walther JH (2017) Iterative Brinkman penalization for simulation of impulsively started flow past a sphere and a circular disc. J Comput Phys 336:261–274

    Article  ADS  MathSciNet  Google Scholar 

  41. Vafai K (1984) Convective flow and heat transfer in variable-porosity media. J Fluid Mech 147:233–259

    Article  ADS  Google Scholar 

  42. Valdés-Parada FJ, Alvarez-Ramírez J, Goyeau B, Ochoa-Tapia JA (2009) Computation of jump coefficients for momentum transfer between a porous medium and a fluid using a closed generalized transfer equation. Transp Porous Media. https://doi.org/10.1007/s11242-009-9370-9

    Article  Google Scholar 

  43. Valdés-Parada FJ, Aguilar-Madera CG, Ochoa-Tapia JA, Goyeau B (2013) Velocity and stress jump conditions between a porous medium and a fluid. Adv Water Resour 62:327–339. https://doi.org/10.1016/j.advwatres.2013.08.008

    Article  ADS  MATH  Google Scholar 

Download references

Acknowledgements

All the numerical simulations have been run on PLAFRIM platform supported by IMB, University of Bordeaux, and INRIA Bordeaux-Sud Ouest.

Funding

There is no funding for this work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Charles-Henri Bruneau.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix 1

The values of the components of the tensor \({{\mathbf {\mathsf{{H}}}}}^{-1}\) are computed for a square pattern of parallel cylinders of square cross section having a porosity \(\epsilon =0.913\) and a flow orthogonal to the cylinders axes. The components \(H_{xx}\), \(H_{yx}\), \(H_{xy}\) and \(H_{yy}\), of tensor \({{\mathbf {\mathsf{{H}}}}}\) are satisfying the following properties:

$$\begin{aligned} H_{xx}(\theta ,Re_{local}) &= H_{yy}(90^\circ -\theta ,Re_{local}),\quad \; H_{yx}(\theta ,Re_{local}) = H_{xy}(90^\circ -\theta ,Re_{local}) \\ H_{xy}(\theta ,Re_{local})&= H_{yx}(90^\circ -\theta ,Re_{local}),\quad \; H_{yy}(\theta ,Re_{local}) = H_{xx}(90^\circ -\theta ,Re_{local}). \end{aligned}$$

Then, for \(\theta \) from \(95^\circ \) to \(180^\circ \) the components of the tensor are given by:

$$\begin{aligned} H_{xx}(\theta ,Re_{local})&= H_{xx}(180^\circ -\theta ,Re_{local}),\quad \; H_{yx}(\theta ,Re_{local}) = -H_{yx}(180^\circ -\theta ,Re_{local}) \\ H_{xy}(\theta ,Re_{local})&= -H_{xy}(180^\circ -\theta ,Re_{local}),\quad \; H_{yy}(\theta ,Re_{local}) = H_{yy}(180^\circ -\theta ,Re_{local}). \end{aligned}$$

Finally, for \(\theta \) from \(185^\circ \) to \(360^\circ \) the components of the tensor are given by:

$$\begin{aligned} H_{xx}(\theta ,Re_{local})&= H_{xx}(360^\circ -\theta ,Re_{local}),\quad \; H_{yx}(\theta ,Re_{local}) = -H_{yx}(360^\circ -\theta ,Re_{local}) \\ H_{xy}(\theta ,Re_{local})&= -H_{xy}(360^\circ -\theta ,Re_{local}), \; H_{yy}(\theta ,Re_{local}) = H_{yy}(360^\circ -\theta ,Re_{local}). \end{aligned}$$

The same symmetries are applicable to the tensor \({{\mathbf {\mathsf{{H}}}}}^{-1}\), the component values of which are provided as additional material.

Appendix 2

In this appendix, a heuristic modification of the K model is explored. The idea is to investigate the possibility of keeping the same form of the effective-medium momentum equation, which has the advantage of being rather simple and solved with the same numerical code, by empirically modifying the Darcy-like penalization term.

An examination of the flow around the front of the porous region with the K model reveals that there is almost no transition between the fluid and the porous medium (see Fig. 10 left) at \(Re=1000\). The zoom of the velocity field corresponds to the same part of the domain shown in Fig. 3. Indeed, for the K model the mean velocity inside the porous domain (averaged with respect to y, which varies from 0 to 30) near the porous medium entrance is quite large. It ranges from 0.83 at \(x=100\) to 0.14 at \(x=200\) (Table 3) instead of 0.33 at \(x=100\) and 0.01 at \(x=200\) for the reference flow. Data reported in Table 3 show that due to the re-circulation zone on top of the porous medium, the mean velocity inside the porous region becomes negative in the second part of this re-circulation zone, namely between \(x=250\) and \(x=300\). However, the velocity is always positive with the K model.

Fig. 10
figure 10

Zoom at the same location of the instantaneous velocity field around the front of the porous rectangle for \(Re=1000\). Left: the K model, right: the improved KU model. Both models are computed on the G6 grid. The length and color of the velocity vectors are related to the velocity magnitude from 0 (dark blue) to the maximum (dark red). (Color figure online)

In this model, the coefficient \(k/\epsilon \) in \({\varvec{\Gamma }}\) is equal to 0.0453 in the present computation. For \(Re=1000\), it yields \({\varvec{\Gamma }}^{-1}=45.3{{\mathbf {\mathsf{{I}}}}}\). This value seems correct in the middle part of the porous medium but does not perform well near the fluid-porous medium boundary, specially when there is a strong normal velocity ahead. To decrease the velocity inside the porous medium close to the surface, it is necessary to decrease locally the value of \({\varvec{\Gamma }}^{-1}\). This can be done, for example, by arbitrarily dividing \({\varvec{\Gamma }}^{-1}\) by an order of magnitude at the boundary and then increasing it linearly in the first \(10\%\) of the rectangle size to reach the right value \({\varvec{\Gamma }}^{-1}=45.3{{\mathbf {\mathsf{{I}}}}}\) and keeping it until the exit section of the porous medium. This can be done by replacing \({\varvec{\Gamma }}^{-1}\) by \({\tilde{\varvec{\Gamma }}}^{-1}\) defined as \({\tilde{\varvec{\Gamma }}}^{-1}=\frac{4}{144-x}{\varvec{\Gamma }}^{-1}\) for \(100\le x\le 140\). The resulting model is called the KK model in the following.

Table 3 Comparison of the mean velocity inside the porous rectangle between the reference flow computed by DNS on the \(\widetilde{G8}\) grid and the three K models computed on the G6 grid at various locations

The solution computed with the KK model has a mean value very close to the reference flow in the entrance region of the porous medium, mostly for \(100\le x\le 200\). This is very encouraging but the flow computed with the KK model is still quite different from the reference flow further into the porous medium (see Fig. 11). An alternative way to modify the penalization coefficient for the effective medium K model is to relate this coefficient to the local velocity magnitude. Then the penalization coefficient can be modulated not only in the front part of the porous medium but within the whole porous medium. With this approach it is possible, in principle, to also encompass the situation when a strong vortex travels down on top of the porous medium. When the velocity magnitude exceeds a chosen threshold, the penalization coefficient is reduced accordingly. Analyzing the velocity inside the porous medium, this can be applied when the velocity magnitude exceeds a value around 0.1. Therefore, a modified value of \({\varvec{\Gamma }}\) can be proposed as \({\tilde{\varvec{\Gamma }}}^{-1}=\frac{1}{100(||{\mathbf {u}}||-0.1)}{\varvec{\Gamma }}^{-1}\) if \(||{\mathbf {u}}||>0.11\). The resulting model is called the KU model in the following.

Fig. 11
figure 11

Mean velocity profiles at three different vertical sections positioned along the x-direction at 200, 400 and 500 for \(Re=1000\). Left: zoom around the front of the porous rectangle, right: zoom in the middle part above the porous rectangle. The two penalization models computed on the G6 grid are compared to the reference flow computed by DNS on the \(\widetilde{G8}\) grid

As shown in Table 3, the KU model does not perform better than the KK model in the front part of the porous medium. However, the predictions are much improved downstream as the KK model yields a very poor approximation (Fig. 11 left). Moreover, the approximation is much better with the KU model in the whole domain at the exit section (Fig. 11 right). The empirical improvements made on the K model, make possible to find a coherent behaviour at the front of the porous body as shown in Fig. 10 right. The porous rectangle is now clearly visible as well as the re-circulation above like in Fig. 3 right for the DNS. As expected the two snapshots can not be directly superimpose because they represent instantaneous solutions that are necessarily different with two approaches so far apart which have a different transient. In addition the Kelvin-Helmholtz vortices are smaller on the finest \(\widetilde{G8}\) grid. However, what is captured by the effective models is the mean flow behaviour.

Fig. 12
figure 12

Mean velocity profiles at three different vertical sections positioned along the x-direction at 200, 400 and 500 for \(Re=1000\). Left: zoom around the front of the porous rectangle, right: zoom in the middle part above the porous rectangle. The KU model computed on the G6 grid is compared to the reference flow computed by DNS on the \(\widetilde{G8}\) grid and to the H model computed on the G7 grid

To conclude this section, it is worth comparing the performance of the KU model computed on the G6 grid and the best one obtained so far, i.e., the H model computed on the G7 grid, with respect to the DNS reference obtained on the \(\widetilde{G8}\) grid for \(Re=1000\). In Fig. 12, the corresponding horizontal velocity profiles, u, are presented. In the porous region (left graph of Fig. 12), the modification on \({\varvec{\Gamma }}\) in the KU model improves significantly the prediction of u at the entrance, although a re-circulation is obtained which does not correspond to the reference flow. The improvement is less significant at the center of the porous region whereas the prediction deteriorates at the exit. Above the porous region, (right part of Fig. 12), the KU model improves the prediction of u in all the three investigated sections, in particular at \(x=200\). The effective integration of the interface, in the case of high speed, to compute the full tensor \({{\mathbf {\mathsf{{H}}}}}\) should provide results close to the DNS. Finally, it must be pointed out that the KU model only requires \(2\%\) of the DNS CPU time.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bruneau, CH., Lasseux, D. & Valdés-Parada, F.J. Comparison between direct numerical simulations and effective models for fluid-porous flows using penalization. Meccanica 55, 1061–1077 (2020). https://doi.org/10.1007/s11012-020-01149-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11012-020-01149-7

Keywords

Navigation